25.02.2015 Views

Rod Failure - Alberta Oil Tool

Rod Failure - Alberta Oil Tool

Rod Failure - Alberta Oil Tool

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A Special Report from <strong>Alberta</strong> <strong>Oil</strong> <strong>Tool</strong>


<strong>Failure</strong> Mechanisms ……………………………….. 1<br />

Design and Operation <strong>Failure</strong>s ………………….. 3<br />

Mechanical <strong>Failure</strong>s ……………………………….. 4<br />

Bent <strong>Rod</strong> <strong>Failure</strong>s ………………………………….. 5<br />

Surface Damage <strong>Failure</strong>s ………………………… 5<br />

Connection <strong>Failure</strong>s ……………………………….. 6<br />

Corrosion <strong>Failure</strong>s …………………………………. 9<br />

Acid Corrosion ……………………………………... 10<br />

Chloride Corrosion ………………………………… 10<br />

CO 2 Corrosion ………………………………….…… 10<br />

H 2 S Corrosion ………………………………………. 11<br />

Microbiologically Influenced Corrosion (MIC) ... 11<br />

Oxygen Enhanced Corrosion ……………………. 12<br />

Scale Corrosion ……………………………………. 12<br />

Stray Current Corrosion ………………………….. 13<br />

Manufacturing Defects ……………………………. 13


Root Cause <strong>Failure</strong> Analysis is Essential for <strong>Failure</strong><br />

Frequency Reduction in Wells With Artificial Lift.<br />

Excerpted from Sucker <strong>Rod</strong> <strong>Failure</strong> Analysis by: Clayton T. Hendricks and Russell D. Stevens<br />

Most failures associated with artificial lift systems can be attributed to one of three downhole<br />

components - pump, sucker rod, or tubing. A pump, sucker rod, or tubing failure is defined as any catastrophic<br />

event requiring servicing personnel to pull or change-out one or more of these components. By this definition,<br />

the failure frequency rate is the total number of component failures occurring per well, per year. Marginally<br />

producing wells with high failure frequency rates are often classified as “problem” wells and effective failure<br />

management practices can mean the difference between operating and plugging these wells. <strong>Failure</strong><br />

management includes preventing, identifying, implementing and recording the “real” root cause of each failure<br />

and is central to overall cost-effective asset management. For the purpose of this photo essay, we will deal only<br />

with sucker rod failures.<br />

Cost-effective failure management begins with prevention, and the time to stop the next failure is nowprior<br />

to an incident! Simply fishing and hanging the well on after a sucker rod failure will not prevent failure<br />

recurrence. In fact, most failures continue with increasing frequency until the entire rod string must be pulled<br />

and replaced. Achievable failure frequency reductions require accurate failure root cause analysis and the<br />

implementation of corrective action measures to prevent failure recurrence. A database capable of querying the<br />

well “servicing” history is needed to tack and identify failure trends. Once a failure trend is identified, remedial<br />

measures should be implemented during well servicing operations to prevent premature rod string failures. The<br />

database failure history should include information on the failure type, location, depth, root cause, and the<br />

corrective action measures implemented.<br />

Sucker rods can be caused to fail prematurely. Understanding the effects of seemingly minor damage<br />

to rod strings, and knowing how that damage can produce catastrophic failures, is very important for production<br />

personnel. Sucker rod failure analysis is challenging and you need to be able to look past the obvious and seek<br />

clues from the not so obvious. All production personnel should have adequate knowledge and training in failure<br />

root cause analysis. Understanding how to identify failures and their contributing factors allows us an<br />

understanding of what is required to correct the root cause of the failure. Every step that can be taken to<br />

eliminate premature sucker rod failures must be taken. On-going training programs concerning sucker rods<br />

should include formal and informal forums that advocate following the recommendations of manufacturers for<br />

artificial lift design, care & handling, storage & transportation, running & rerunning, and makeup & breakout<br />

procedures. A variety of training schools are currently available and, with advanced notice, most can be tailored<br />

to meet the specific needs of production personnel.<br />

<strong>Failure</strong> Mechanisms<br />

All sucker rod, pony rod, and coupling fractures are either tensile or fatigue failures. Tensile failures<br />

occur when the applied load exceeds the tensile strength of the rod. The load will concentrate at some point in<br />

the rod string, create a necked-down appearance around the circumference of the rod, and a fracture occurs<br />

where the cross-section is reduced. This rare failure mechanism only occurs when pulling too much load on the<br />

rod string- such as attempting to unseat a stuck pump. To avoid tensile failures, the maximum weight indicator<br />

pull for a rod string in “like new” condition should never exceed 90% of the yield strength for the known size and<br />

grade of the smallest diameter sucker rod. For unknown sucker rod conditions, sizes, or grades a sufficient derating<br />

factor should be applied to the maximum weight pulled. All other sucker rod, pony rod and coupling<br />

failures are fatigue failures.<br />

Fatigue failures are progressive and begin as small stress cracks that grow under the action of cyclic<br />

stresses. The stresses associated with this failure have a maximum value that is less than the tensile strength<br />

of the sucker rod steel. Since the applied load is distributed nearly equally over the full cross-sectional area of<br />

the rod string, any damage that reduces the cross-sectional area will increase the load or stress at that point<br />

and is a stress raiser. A small stress fatigue crack forms at the base of the stress raiser and propagates<br />

perpendicular to the line of stress, or axis of the rod body. As the stress fatigue crack gradually advances, the<br />

1


mating fracture surfaces opposite the advancing crack front try to separate under load and these surfaces<br />

become smooth and polished from chafing. As the fatigue crack progresses, it reduces the effective crosssectional<br />

area of the sucker rod until not enough metal remains to support the load, and the sucker rod simply<br />

fractures in two. The fracture surfaces of a typical fatigue failure have a fatigue portion, tensile portion, and final<br />

shear tear.<br />

Fatigue failures are initiated by a multitude of stress raisers. Stress raisers are visible or microscopic<br />

discontinuities that cause an increase in local stress on the rod string during load. Typical visible stress raisers<br />

on sucker rods, pony rods and couplings are bends, corrosion, cracks, mechanical damage, threads, and wear<br />

or any combination of the preceding. This increased stress effect is the most critical when the discontinuity on<br />

the rod string is transverse (normal) to the principle tensile stress. In determining the stress raiser of a fatigue<br />

failure, the fatigue portion opposite the final shear tear (extrusion/protrusion) must be carefully cleaned and<br />

thoroughly examined. Fatigue failures have visible or macroscopic identifying characteristics on the fracture<br />

surface, which help to identify the location of the stress raiser. Ratchet marks and beach marks are arguably<br />

two of the most important features in fatigue failure identification. Ratchet marks are lines that result from the<br />

intersection and connection of multiple stress fatigue cracks while beach marks indicate a change in the rate of<br />

crack growth. Ratchet marks are parallel to the overall direction of crack growth and lead to the initiation point<br />

of the failure. Beach marks are elliptical or semi-elliptical rings radiating outward from the fracture origin and<br />

indicate successive positions of the advancing stress fatigue crack growth.<br />

Figure 1<br />

Figure 1 is an example of tensile and fatigue<br />

failures. The two examples on the right are tensile<br />

failures. A tensile failure is characterized by a reduction<br />

in the diameter of the cross-sectional area at the point of<br />

fracture. Typical tensile failures have cup-cone fracture<br />

halves. The second example from the right is typical in<br />

appearance for tensile failures. During the final stages of<br />

an overload, fractures from tensile failures rupture, or<br />

shear, on 45ϒ angles to the stresses applied. A good<br />

example of the shear is the characteristic cup-cone<br />

fracture surfaces of a typical tensile failure. The rod body<br />

on the right is an excellent example of needing to look<br />

past the obvious for the not so obvious. A fatigue failure<br />

is primarily responsible for the failure even though fracture occurred while trying to unseat a stuck pump. Visual<br />

examination of the fracture surface reveals a small semi-elliptical, stress fatigue crack. This sucker rod had preexisting,<br />

transverse fatigue cracks, from in-service stresses. One of the stress fatigue cracks opened during the<br />

straight, steady load applied in attempting to unseat the pump, and fracture occurred. The tensile failure is<br />

secondary and results in the unusual appearance of the fracture surface – with the small fatigue portion, large<br />

tensile portion and unusually large 45ϒ double shear tears.<br />

The remaining examples are fatigue failures on: casehardened sucker rods; normalized and tempered<br />

sucker rods; and quenched and tempered sucker rods. The example on the far left is a torsional fatigue failure<br />

from a progressing cavity pump. Ratchet marks found in the large fatigue portion, and originating from the<br />

surface of the rod body, completely encircle the fracture surface with the small tensile tear portion shown slightly<br />

off middle-center. The second rod body on the left is a casehardened fatigue failure. The case encircling the<br />

rod body diameter carries the load for this high tensile strength sucker rod and if you penetrate the case, you<br />

effectively destroy the load-carrying capability of this type of manufactured sucker rod. The fatigue crack<br />

advances around the case and progresses across the rod body. A fatigue failure on a casehardened sucker rod<br />

generally exhibits a small fatigue portion and a large tensile tear. The third rod body from the left is typical in<br />

appearance for most fatigue failures. Typical fatigue failures have a fatigue portion, tensile portion with a final<br />

shear tear. The width of the fatigue portion is an indication of the loading involved with the fracture. Mechanical<br />

damage can prevent or hinder failure analysis by destroying the visual clues and identifying characteristics<br />

normally found on a fatigue fracture surface. Care must be exercised when handling the fracture halves. It is<br />

very important to resist the temptation to fit the mating fracture surfaces together since this almost always<br />

destroys (smears) microscopic features. To avoid mechanical damage, fracture surfaces should never actually<br />

touch during fracture-surface matching.<br />

2


Design and Operation <strong>Failure</strong>s<br />

Sucker rod failure prevention begins with design. It is possible for poorly designed rod strings to<br />

contribute to other component failures in the artificial lift system, such as rod cut tubing resulting from<br />

compressive rod loads. Designing the artificial lift system is a compromise between the amount of work to be<br />

done and the expense of doing this work over a cost-effective period of time. Numerous combinations of<br />

depths, tubing sizes, fluid volumes, pump sizes and configurations, unit sizes and geometries, stoke lengths,<br />

pumping speeds and rod tapers are available to the system designer. Sucker rod size and grade selection is<br />

dependent upon many factors including predicted maximum stresses, stress ranges, and operating<br />

environments.<br />

Figure 2 Commercially available computer design<br />

programs allow the system designer to optimize<br />

production equipment at the least expense for the well<br />

conditions existing at the time of the design. After the<br />

initial design and installation of the rod string, periodic<br />

dynamometer surveys should be utilized to confirm that<br />

equipment load parameters are within those considered<br />

acceptable. A good initial design may become a poor<br />

design if well conditions change. Changes in the fluid<br />

volume, fluid level, stroke length, stokes per minute or<br />

pump size severely impact the total artificial lift system.<br />

Changes in fluid corrosiveness can affect the fatigue<br />

endurance lift of sucker rods and may lead to premature<br />

failures. When one of the preceding conditions change, the design of the artificial lift system must be reevaluated.<br />

Figures 2 and 3 are examples of design and<br />

operationally induced mechanical failures. Wear, flexing<br />

fatigue, unidirectional bending fatigue, and stress-fatigue<br />

failures indicate compressive rod loads, deviated wells,<br />

fluid pound, gas interference, highly stressed sucker rods,<br />

improperly anchored tubing, pumps tagging bottom,<br />

sticking pump plungers, unanchored tubing, or some<br />

combination of the preceding.<br />

Wear causes rod failures by reducing the crosssection<br />

of the metal, exposing new surface metal to<br />

Figure 3 corrosion, and causes joint failures from impact and<br />

shoulder damage. The Class T coupling on the left, the<br />

Class SM coupling second from left, and the rod body on the left are all examples of wear. Wear on the sucker<br />

rod string is defined as the progressive removal of surface metal by contact with the tubing. Wear that is equal<br />

in length, width, and depth usually suggests a deviated or crooked well bore. Angled wear patterns indicate rod<br />

strings that are aggressively contacting the tubing at an angle, usually as a result of fluid pound or unanchored<br />

(improperly anchored) tubing. The middle rod body represents corrosion-abrasion wear. Wear also removes<br />

corrosion inhibiting films and exposes new surface metals to corrosive well fluids-which accelerate the rate of<br />

corrosion. The Class T coupling on the far right has a work-hardened ridge from tubing-slap wear. Tubing-slap<br />

wear is the result of the rod string “stacking out” – probably as a result of fluid pound, gas interference or pump<br />

tagging. The work-hardened material doesn’t wear as fast as the softer material on either side of the workhardened<br />

area, and it leaves a ridge of material as the rest of the coupling wears.<br />

The second rod body from the left is a flexing fatigue failure. Flexing fatigue failures occur from the<br />

motion of the rod string having a constant lateral or side movement during the pumping cycle. Stress fatigue<br />

cracks due to flexing will concentrate along the area of the rod where the greatest bending stresses occurred.<br />

The transverse, stress fatigue cracks will be on one half of the circumference of the rod body, closely spaced<br />

near the rod upsets and gradually spreading apart moving toward the middle of the rod body. Most flexing<br />

fatigue failures occur above the connection in the transition zone of the rod body-between the rigid coupling and<br />

upset area and the more flexible rod body. Flexing fatigue failures will not show permanent bends since this<br />

problem occurs while the rod string is in motion. The example on the far right is a unidirectional bending fatigue<br />

3


failure. This type of failure generally has two tips protruding above the fracture surface. These distinct failure<br />

characteristics indicate a double shear-lip tear. Double shear-lip tears are the direct result of unidirectional<br />

bending stresses, with fractures and the fatigue damage occurring under compressive rod loads. Compressive<br />

rod loads may be the result of large bore pumps with small diameter sucker rods or multiple tapers in shallow<br />

wells.<br />

The second rod body sample on the right (Figure 3) is a stress fatigue failure. Stress fatigue failures<br />

occur on highly stressed sucker rods as a result or worn out sucker rods, overloads, or extremely high rod loads<br />

for short periods of time. Stress fatigue failures may have closely spaces, fine, transverse secondary fatigue<br />

cracks that completely encircle the circumference of the rod body. The stress fatigue cracks may be on the<br />

wrench square and over the entire length of the rod body. With very old sucker rods, stress fatigue cracks and<br />

failure may occur within normal everyday operating loads.<br />

Figure 4 is an example of coupling-to-tubing slap.<br />

Coupling-to-tubing slap is the result of extremely<br />

aggressive angle contact to the tubing by the rod string.<br />

This aggressive contact is the direct result of severe fluid<br />

pound, unanchored (or improperly anchored) tubing,<br />

sticking (or stuck) pump plungers, or any combination of<br />

the preceding.<br />

Figure 5<br />

Figure 4<br />

Figure 5 is an example of rod guide related<br />

damage. The example on the left is a reconditioned, high<br />

tensile strength sucker rod. Turbulent fluid flow,<br />

associated with short, blunt-end injection molded rod<br />

guides, allowed crevice corrosion in the critical wash area<br />

around the end of the guide. Prior to inspecting the moldon<br />

rod guides were removed from the rod body for<br />

reconditioning. Class 1 reconditioned sucker rods cannot have discontinuities greater than 20 mils (0.020”) per<br />

API Specification 11BR. The crevice corrosion was under the 20 mils allowed for a Class 1 reconditioned<br />

sucker rod. However, the notch sensitivity (discontinuity intolerance) of a high tensile strength sucker rod is<br />

high. In other words, small pits can be detrimental to the high tensile stresses associated with the high strength<br />

sucker rod and reconditioned high strength sucker rods should be de-rated for load. The example in the middle<br />

is an erosion/corrosion failure resulting from short, blunt-end; field applied rod guides in small tubing with high<br />

fluid velocities. Erosion/corrosion pits will be “fluid cut” with very smooth bottoms. Pit shape characteristics may<br />

include sharp edges and steep sides if accompanied by CO 2 or broad smooth pits with beveled edges if<br />

accompanied by H 2 S. The example on the right is abrasion wear from a field-applied guide moving up and<br />

down on the rod body during the pumping cycle. Generally speaking, mold-on rod guides provide better laminar<br />

flow, a minimum of three to four times more bonding and holding power and are more cost-effective than are<br />

field applied rod guides.<br />

Mechanical <strong>Failure</strong>s<br />

Mechanical failures account for a large percentage of the total number of all rod string failures.<br />

Mechanical failures include every type except manufacturing defects and stress/corrosion fatigue. Mechanical<br />

damage to the rod string contributes to a stress raiser which will cause sucker rod failures. The time to failure<br />

will be influenced by many variables, of which maximum stress, operating environment, orientation of the<br />

damage, sucker rod chemistry, sucker rod heat treatment type, stress range and type of damage will be of the<br />

most important. Mechanical damage can be caused by inept artificial lift design, improper care and handling<br />

procedures, careless makeup and breakout procedures, out-of-date operating practices, or any combination of<br />

these elements.<br />

4


Bent <strong>Rod</strong> <strong>Failure</strong>s<br />

Bending fatigue failures account for a significant number of all mechanical failures. It is a fact that all<br />

bent sucker rods eventually fail. New sucker rods are manufactured to a body straightness of no less than 1/16<br />

inch in any twelve inches of rod body length. Sucker rods within this tolerance of straightness will roll easily on<br />

a level rack with five supports. Any degree of bend greater than this will cause an increase in local stress at the<br />

point of the bend during applied load. When the bent rod body is pulled straight during load, the fatigue strength<br />

of the material is quickly reached. The cycle of continually exceeding the ultimate material strength is repeated<br />

during the pumping cycle and causes stress fatigue cracks on the concave side of the bend. These stress<br />

fatigue cracks progress across the bar, during load, until not enough metal remains in the bar to support the<br />

load, and fracture occurs.<br />

Straightening the raw bar stock is the first step in the process of manufacturing sucker rods. Cold<br />

straightening the bar deforms the grain structure below its recrystallization temperature, putting a strain in the<br />

bar that is accompanied by a work hardening effect. During the manufacturing process, the function of heat<br />

treatment is to stress-relieve the residual and induced stresses caused by bar rolling, bar straightening<br />

processes and from forging the rod upsets. Heat treatment changes the metallurgical structure of the forged<br />

ends to match that of the rod body and also controls the mechanical properties of the sucker rod. Any rod body<br />

bend created after hear treatment causes work hardening, which creates an area of hardness different than the<br />

surrounding surfaces. This condition is referred to as a “hard spot” and is a stress raiser to load. Mechanical<br />

processing, such as passing the finished sucker rod through a system of rollers, will attempt to remove the bend<br />

so it appears to be straight. However, reconditioning processes are not capable of stress relieving bent sucker<br />

rods. A bent sucker rod is permanently damaged and should not be used because all bent sucker rods will<br />

eventually fail.<br />

Figure 6 (with inset of Figure 7) is an example of<br />

binding fatigue failures. Bending fatigue failures can be<br />

identified by the angled fracture surface, which will be at<br />

some angle other than 90ϒ to the axis of the rod body.<br />

The example on the left illustrates a fracture caused by a<br />

long radius bend, or gradual bow in the rod body (left<br />

example in Figure 7). The fracture surface shows fatigue<br />

damage, and has a slight angle when compared to the<br />

axis of the rod body. The middle example is a short<br />

radius bend (right example in Figure 7). The fracture<br />

surface is at a greater angle to the axis of the rod body<br />

with a small fatigue portion and a large tensile tear<br />

portion. The example on the right is the result of a<br />

Figure 6 Figure 7<br />

corkscrewed sucker rod. Notice how convoluted the fracture surface is in<br />

appearance. As a general rule, the greater the bend in the rod body, the more<br />

convoluted the fracture surfaces appear. In operation, the time for the rod to fracture<br />

is greatly shortened. Poor care and handling procedures usually cause bent rods.<br />

Surface Damage <strong>Failure</strong>s<br />

Everything possible should be done to prevent mechanical surface damage<br />

to sucker rods, pony rods and couplings. Surface damage increases stress during<br />

applied loads, potentially causing rod string failures. The type of damage, and its orientation, contributes to this<br />

increased stress effect. The orientation of the damage contributes to higher stresses with transverse damage<br />

having increased stresses over those associated with longitudinal damage. A sharp nick will create a higher<br />

stress concentration and would be more detrimental to load than a shallow, broad-based depression. Sucker<br />

rods with indications of surface damage must not be used and must be replaced. Care should be used to avoid<br />

all metal-to-metal contact that might result in dents, nicks, or scratches. To prevent potential sucker rod<br />

damage, place strips of wood between metal storage racks and between each layer of sucker rods so metal-tometal<br />

contact can be avoided. Use sucker rods for what they were designed for- to lift a load. Never use sucker<br />

rods as a walkway or workbench. Keep metal tools not intended for use on sucker rods and all other metal<br />

objects away from the rods. Make sure the tool you use is intended for the purpose and ensure that it is in<br />

proper working order.<br />

5


Figure 8 is an example of various surface<br />

damage failures. The example on the left shows a slight<br />

depression from a wrench, tool, or other metal object.<br />

The second example from the left is damage from a pipe<br />

wrench used in applying field-installed rod guides. The<br />

second example from the right has a small longitudinal<br />

scratch, through metal-to-metal contact, by allowing<br />

sucker rods to run down other rods in a rod bundle during<br />

installation. The example on the right exhibits transverse<br />

surface damage.<br />

Figure 8<br />

Figure 9 is an example of surface damage<br />

caused by sucker rod elevators. The bottom example is<br />

damage from worn or misaligned elevator seats. After an extended period of service, the elevator seats<br />

become so worn that they develop an oval shape rather than a round shape. As the oval shape grows, the<br />

tangency ring of the rod upset to the elevator seat face is<br />

lowered in the front half of the seat. As the seat continues<br />

to wear the seating position of the rod upset is moved<br />

forward of the elevator trunnion centerline. This causes<br />

an offset in the hook load and tilts the elevator body<br />

forward. When the elevator lifts the rod string load, the<br />

hook load will bend the sucker rod centerline to coincide<br />

with the elevator trunnion centerline. As the rod string<br />

weight increases, the hook load will bend every sucker<br />

rod engaged by this elevator. Bent sucker rod failures<br />

that occur below the surface upset bead may be from bad<br />

elevator seats. The top example is damage caused by<br />

the elevator latches. This type of damage normally<br />

occurs as a result of picking up or laying down in doubles.<br />

Never pick up or lay down anything more than one single sucker rod. Anything else causes the elevator latches<br />

to act as a fulcrum and allow bending stresses to concentrate in the transition zone of the rod body and the<br />

forged upset.<br />

Connection <strong>Failure</strong>s<br />

Figure 9<br />

The API sucker rod connection is designed as a shouldered, friction loaded connection. Since the<br />

fatigue endurance of the sucker rod connection is low when subjected to cyclic loads it is necessary to limit the<br />

cyclic loads with pin preload. If the pin preload is greater than the applied load the load in the connection<br />

remains constant and no fatigue occur from cyclic loads. The friction load that develops between the pin<br />

shoulder face and the coupling shoulder face, helps lock the connection together to prevent it from coming<br />

unscrewed downhole. However, if the preload is less than the applied load, the pin shoulder face and the<br />

coupling shoulder face will separate during the cyclic motion of the pumping unit. Once these faces separate<br />

the connection is cyclically loaded and will result in a loss of displacement, or loss of tightness, failure. Loss of<br />

displacement failures may result from improper lubrication, inadequate makeup, over-torque, tubing-slap wear,<br />

or any combination of these elements.<br />

Figure 10 is an example of pin failures due to a loss<br />

of displacement. The sample on the right is typical in<br />

appearance for a loss of displacement pin failure.<br />

Insufficient makeup, or the loss of tightness, caused the pin<br />

shoulder face and the coupling shoulder face to separate.<br />

When these faces separate, a bending moment is added to<br />

the tensile load in the pin. The threaded section of the pin<br />

is held rigid while the rest of the pin flexes. The motion of<br />

the rod string causes stress fatigue cracking to start in the<br />

first fully formed thread root above the undercut. Small<br />

fatigue cracks begin along the thread root and consolidate<br />

into a major stress fatigue crack. The fracture surface of a<br />

typical loss of displacement pin failure has a small fatigue<br />

Figure 10<br />

6


portion covering approximately one-third of the fracture surface with the tensile tear portion and final shear tear<br />

covering the remaining fracture surface. The examples on the left and in the middle will occur as a result of<br />

stress loading when stress-raising factors such as corrosion or mechanical damage is present on the surface of<br />

the pin undercut.<br />

Figure 11 is another example of two types of pin<br />

failures. The sample on the left is typical in appearance<br />

for a loss of displacement pin failure. However, this pin<br />

fracture was caused by the hydraulic rod tongs during<br />

makeup as is evidenced by the stair-stepped tensile tear.<br />

It is not uncommon for pin fractures to occur at makeup, if<br />

the pin has a pre-existing stress fatigue crack due to the<br />

high torque required during joint makeup, with large<br />

diameter Class D and all sizes of high tensile strength<br />

sucker rods. The sample on the right is an example of<br />

excessive torque on a soft pin. The fracture surface has a<br />

large fatigue portion, with multiple ratchet marks in the<br />

pin-thread root, and a small tensile portion.<br />

Figure 12 is an example of a loss of displacement<br />

coupling failure. The fracture initiated in the coupling<br />

thread-root opposite the first fully formed pin starting<br />

thread. One-third/two-third fracture halves, in length, with<br />

ratchet marks originating in the thread root indicate a loss<br />

of displacement coupling failure. The fracture surface of a<br />

typical loss of displacement coupling failure has a small<br />

fatigue portion and a large tensile tear portion. Loss of<br />

displacement coupling failures are primarily associated<br />

with Class D sucker rods and high tensile strength sucker<br />

rods.<br />

Figure 11<br />

Figure 12<br />

Mid-length coupling fractures, with ratchet marks<br />

leading from the outside, indicate another type of failure. The fatigue crack starts from the outside coupling<br />

surface, progressing inward toward the threads, then around the coupling wall to a tensile fracture. Mid-length<br />

fractures indicate coupling failures from mechanical damage to the coupling surface, exceeding the stress<br />

fatigue endurance limit of the material, or a manufacturing defect. Most mid-length coupling fractures are the<br />

result of mechanical damage or over load. Mid-length coupling fractures due to overload have a small fatigue<br />

portion and large tensile tear portion. This failure is common with high strength sucker rods and Class SM<br />

couplings. Use Class T couplings to avoid mid-length coupling failures with high tensile strength sucker rods.<br />

Figure 13 is an example of thread galling in the<br />

sucker rod connection. Thread galling is mechanical<br />

damage to the sucker rod and/or coupling threads.<br />

Thread galling is the result of damaged or contaminated<br />

threads causing the interference between the threads to<br />

be great enough to rip and tear the thread surfaces. The<br />

threads weld together during makeup and strip apart at<br />

breakout and the connection is damaged and destroyed<br />

beyond use. Hard stabbing damage to the leading<br />

thread, contaminated threads and cross threading on<br />

larger diameter rods are the primary causes of thread<br />

galling. Cleaning the threads prior to makeup, properly<br />

lubricating the threads and following careful makeup<br />

procedures will prevent thread galling.<br />

Figure 13<br />

Figure 14 is an example of wrench square failures. Wrench square failures are extremely rare and<br />

seldom occur unless from mechanical damage, corrosion, manufacturing defects or torsional stress causing<br />

fatigue damage. The example on the left is a wrench square failure from severe mechanical damage. A loose<br />

7


or sloppy backup on the hydraulic rod tongs has rounded<br />

the wrench square corner. The stress fatigue crack<br />

began in the corner of the wrench square and progressed<br />

to final rupture or fracture.<br />

Figure 14<br />

The example on the right is a wrench square<br />

failure from a manufacturing defect. The failure initiated<br />

in the die stamp mark and is an example of an excessive<br />

die stamp depth failure. Die stamp markings can become<br />

notches that serve as stress raisers if the depth of the die<br />

stamping, during the forging process, is not controlled and<br />

kept within API Specification 11B, allowable tolerances.<br />

Figure 15 is an example of the damage that<br />

occurs as a result of severely over-tightening the sucker<br />

rod connection. The example shown is an over-tightened<br />

coupling that has flared out or bulged near the contact<br />

face. Slim-hole couplings are more susceptible to this<br />

type of over-tightened damage than are full sized<br />

couplings. Over-tightened full size couplings on Class D<br />

and high strength sucker rods generally exhibit slight<br />

bulges and have the concentric deformation ridge of<br />

material on the coupling shoulder face from the<br />

impression of the pin shoulder face. Over-tightening with<br />

Figure 15<br />

hydraulic rod tongs will twist off soft pins resulting in a<br />

tensile failure appearance. The pin undercut will neck<br />

down and fracture occurs rapidly. With Class D sucker rods, an indication of over-tightening is the concentric<br />

deformation ridge of material on the pin shoulder face from the impression of the coupling shoulder face. Overtightening<br />

on normalized and tempered high tensile strength sucker rods will begin to pull the threads out of the<br />

coupling.<br />

Figure 16<br />

Figure 17<br />

Figure 16 is an example of impact cracks on<br />

couplings. The practice of “warming up,” or hammering,<br />

on couplings in order to loosen them should not be<br />

allowed. This example shows how impact damage to a<br />

Class T coupling causes stress fatigue cracks around the<br />

impact points and accelerated localized corrosion.<br />

Hammering on Class SM couplings causes stress fatigue<br />

cracks in the hard spray surface and results in a coupling<br />

failure due to stress/corrosion fatigue.<br />

Figure 17 is an example of polished rod failures.<br />

The majority of all polished rod failures occur either in the<br />

body, just below the polished rod clamp, or in the pin.<br />

Polished rod body failures below the polished rod clamp<br />

result from the addition of bending stresses. These<br />

bending stresses may be imposed by pumping units out<br />

of alignment, carrier bars that do not set level, worn<br />

carrier bars, misaligned load cells, or incorrect polished<br />

rod clamp installation. The polished rod failure on the left<br />

is an example of polished rod clamp on the sprayed<br />

portion of a Spraymetal polished rod. Spraymetal<br />

polished rods have an unsprayed portion for polished rod<br />

clamp placement. Never put a polished rod clamp on the<br />

sprayed portion of a Spraymetal polished rod. The<br />

polished rod failure on the right has small, longitudinal<br />

scratches caused from mishandling.<br />

8


Polished rod pin failures generally occur due to the installation of sucker rod couplings. Polished rod<br />

pins have a 9ϒ thread taper between the straight-threaded section and the shoulder. Sucker rod couplings have<br />

a 30ϒ starting thread and a deep recess that doesn’t engage all the polished rod pin threads. Polished rod<br />

couplings have a 9ϒ starting thread and a profile designed to properly fit the polished rod pin. The shallow<br />

recess to the first thread easily distinguishes polished rod couplings from sucker rod couplings and allows every<br />

polished rod pin thread to be engaged.<br />

Corrosion <strong>Failure</strong>s<br />

Corrosion is one of the greatest problems encountered with produced fluids and accounts for about onehalf<br />

of all sucker rod failures. Corrosion is the destructive result of an electrochemical reaction between the<br />

steel used in making sucker rods and the operating environment to which it is subjected. Simply put, corrosion<br />

is nature’s way of reverting a man-made material of a higher energy state (steel), back to its basic condition<br />

(native ore) as it is found in nature. The elemental iron in steel combines with moisture or acids, to form other<br />

compounds such as iron oxide, sulfide, carbonate, etc. Some form and concentration of water is present in all<br />

wells considered corrosive and most contain considerable quantities of dissolved impurities and gases. For<br />

instance, carbon dioxide (CO 2 ) and hydrogen sulfide (H 2 S) acid gases, common in most wells, are highly soluble<br />

and readily dissolve in water, which tends to lower its pH. The corrosivity of the water is a function of the<br />

amount of these two gases that are held in solution. All waters with low pH values are considered corrosive to<br />

steel, with lower values representing greater acidity, or corrosiveness.<br />

All downhole environments are corrosive to some degree. Some corrosive fluids may be considered<br />

non-corrosive if the corrosion penetration rate, recorded as mils of thickness lost per year (mpy), is low enough<br />

that it will not cause problems. However, most producing wells are plagued by corrosion problems and no<br />

currently manufactured sucker rod can successfully withstand the effects of this corrosion alone. While<br />

corrosion cannot be completely eliminated it is possible to control its reaction. All grades of sucker rods must be<br />

adequately protected through the use of effective chemical inhibition programs (reference current editions of API<br />

Specification 11BR and NACE Standard RPO195). Some sucker rod grades, due to different combinations of<br />

alloying elements, microstructures and hardness levels, are capable of longer service life in inhibited corrosive<br />

wells than other grades of either low or high tensile strength.<br />

Why do new sucker rods seem to corrode faster than older rods in the same string? Two sucker rods<br />

with the same chemical analysis will form a galvanic corrosion cell if the physical condition of one is different<br />

from the other. Physical differences in a sucker rod may be caused from poor care and handling practices (i.e.<br />

surface damage resulting in bruises, nicks, bends) and/or corrosion deposits. Since new sucker rods go into the<br />

well without corrosion deposits, they often corrode preferentially in relation to rods that are coated with corrosion<br />

deposits. Corrosion on steel starts very aggressively but often slows down as soon as an obstructive surface<br />

film of corrosion deposit (scale) is formed upon the metal surface. For example, CO 2 generates iron carbonate<br />

scale as a by-product of its corrosion. This scale coats the sucker rod and retards the corrosion penetration<br />

rate, which tends to slow down corrosion. However, if this deposit is continuously cracked by a bending<br />

movement or removed by abrasion, aggressive local corrosion continues on the area with the scale removed,<br />

and results in deep corrosion pitting.<br />

Can high tensile strength sucker rods be used in a corrosive environment? Generally soft rods tolerate<br />

corrosion better than hard rods and, as a rule of thumb; you should always use the softest rod that will handle<br />

the load. However, if load requirements dictate the use of high tensile strength rods than it is important to<br />

protect the rods with an effective surface film of corrosion inhibitor. In most cases, if you can adequately protect<br />

downhole equipment from corrosion, you should be able to adequately protect high tensile strength rods from<br />

corrosion by increasing the application frequency of the corrosion-inhibitor program. In other words, if you<br />

effectively batch treat once a week with 40 parts per million (ppm) of corrosion inhibitor for D class rods, you will<br />

need to batch treat twice weekly with 40 ppm of corrosion inhibitor for high tensile strength rods. Treatment<br />

volumes vary and are dependent upon many factors too numerous to list. Always consult with a corrosion<br />

control specialist prior to the installation of every rod string, especially when corrosion fatigue is suspected as<br />

the failure root cause.<br />

9


Figure 18<br />

Figure 18 is an example of corrosion fatigue from<br />

CO 2 corrosion. The size of the pit, as far as when it<br />

becomes detrimental to the rod, depends on two factorsmaterial<br />

type and hardness. Class K sucker rods may<br />

develop deeper and larger pits than a Class D sucker rod<br />

before it becomes detrimental to the rods. Class D sucker<br />

rods may develop deeper and large pits than a high<br />

tensile strength rod before it becomes detrimental to the<br />

rods. Softer materials with lower rod stress tolerate larger<br />

pits than do harder materials with higher rod stress.<br />

Therefore, small pits can be detrimental to higher tensile<br />

strength sucker rods as opposed to a softer rod that does<br />

not have as much rod stress.<br />

Figure 5<br />

Acid Corrosion<br />

Service companies use acids for well<br />

stimulation and cleanout work. All acid work should<br />

have an inhibitor mixed with the acid prior to injection<br />

into the well. Spent acids are still corrosive to steel and<br />

the well should be “flushed“ long enough to recover all<br />

acid. In rare instances, some produced waters contain<br />

organic acids that have formed downhole, such as<br />

acetic, hydrochloric and sulfuric acids. Corrosion from<br />

acid is a general thinning of metal, leaving the surface<br />

with the appearance of sharp, feathery or web-like<br />

residual metal nodules. Metallic scale will not be formed<br />

in the pits. The left sample in Figure 5 is an example of<br />

acid corrosion.<br />

Chloride Corrosion<br />

Chlorides contribute to the likelihood of an increase in corrosion related sucker rod failures. The<br />

corrosivity of water increases as the concentration of chlorides increase. Corrosion inhibitors have more<br />

difficulty reaching and protecting the steel surface of sucker rods in wells with high concentrations of chlorides.<br />

Corrosion, from waters with high concentrations of chlorides, has the tendency to be more aggressive to carbon<br />

steel sucker rods than to alloy steel sucker rods. Chloride corrosion tends to evenly pit the entire surface area<br />

of the sucker rod with shallow, flat-bottomed, irregular shaped pits. Pit shape characteristics include steep walls<br />

and sharp pit edges.<br />

CO 2 Corrosion<br />

CO 2 combines with water to form carbonic acid, which lowers the pH of the water. Carbonic acid is very<br />

aggressive to steel and results in large areas of rapid metal loss that can completely erode sucker rods and<br />

couplings. The corrosion severity increases with increasing CO 2 partial pressure and temperature. CO 2<br />

corrosion pits are round based, deep with steep walls and sharp edges. The pitting is usually interconnected in<br />

long lines but will occasionally be singular and isolated. The pit bases will be filled with iron carbonate scale, a<br />

loosely adhering gray deposit generated from CO 2 .<br />

Figures 19 and 20 show typical examples of CO 2 corrosion. Figure 19 is an example of CO 2 corrosion<br />

on couplings and Figure 20 is an example of CO 2 corrosion on rod bodies.<br />

10


Figure 19<br />

Figure 20<br />

H 2 S Corrosion<br />

Figure 21<br />

H 2 S pitting is round based, deep with steep walls<br />

and beveled edges. It is usually small, random, and<br />

scattered over the entire surface of the rod. A second<br />

corrodent generated by H 2 S is iron sulfide scale. The<br />

surfaces of both the sucker rod and the pit will be covered<br />

with the tightly adhering black scale. Iron sulfide scale is<br />

highly insoluble and cathodic to steel which tends to<br />

accelerate corrosion penetration rates. A third corroding<br />

mechanism is hydrogen embrittlement, which causes the<br />

fracture surface to have a brittle or granular appearance.<br />

A crack initiation point may or may not be visible and a<br />

fatigue portion may not be present on the fracture surface.<br />

The shear tear of a hydrogen embrittlement failure is<br />

immediate during fracture due to the absorption of<br />

hydrogen and the loss of ductility in the steel. Although a<br />

relatively weak acid, any measurable trace amount of H 2 S<br />

is considered justification for chemical inhibition programs<br />

when any measurable trace amount of water is also<br />

present.<br />

Figure 22<br />

Figure 21 and 22 are examples of H 2 S corrosion. The<br />

three rod body samples on the left are examples of<br />

localized corrosion (pitting) and the two rod body samples<br />

on the right are examples of general thinning corrosion<br />

from under-scale deposit corrosion. The sample in Figure<br />

22 is an example of a pin failure due to hydrogen<br />

embrittlement.<br />

Microbiologically Influenced Corrosion (MIC)<br />

Some amount of microscopic life form is present<br />

in essentially every producing well. Of primary concern to<br />

sucker rods are the single celled organisms capable of<br />

living in all sorts of conditions and multiplying with<br />

incredible speed-commonly referred to as bacteria or "bugs”. Suspect fluids should be monitored continuously<br />

for bacteria by sampling, identifying and counting the bacteria. The extinction dilution technique is commonly<br />

used to culture bacteria for an estimation of the number of bacteria present in the well. Bactericide should be<br />

11


used on all suspect fluids to control bacteria populations. Bacteria are classified according to their oxygen<br />

requirements: aerobic (requires oxygen), anaerobic (no oxygen), and facultative (either). Some bacteria<br />

generate H 2 S, produce organic acids or enzymes, oxidize soluble iron in produced waters, or any combination of<br />

the preceding. MIC has the same basic pit shape characteristics of H 2 S, often with multiple stress cracks in the<br />

pit base, tunneling around the pit edge and/or unusual<br />

anomalies (i.e. shiny splotches) on the rod surface. Figure 23<br />

Bacteria are very aggressive and all sucker rod grades<br />

corrode rapidly in downhole environments containing<br />

bacteria. Sulfate reducers (SRB’s), those that produce<br />

H 2 S, probably cause more problems to downhole artificial<br />

lift equipment than do any other bacteria type. Multiple<br />

cracking in the pit bases results from the hydrogen sulfide<br />

by-product of the bacterial lifestyle, which corrode and<br />

embrittle the surface of the steel under the colony. Figure<br />

23 show several examples of microbiologically influenced<br />

corrosion (bacteria) on sucker rod bodies.<br />

Oxygen Enhanced Corrosion<br />

Oxygen enhanced corrosion will be most<br />

prevalent on couplings, with a few instances found on rod<br />

upsets. Oxygen enhanced corrosion is rarely seen on the<br />

rod body. In fact, aggressive oxygen enhanced corrosion<br />

can erode couplings without harming the sucker rods on<br />

either side. The rate of oxygen enhanced corrosion is<br />

directly proportional to the dissolved oxygen<br />

concentration, chloride content of the produced water<br />

and/or presence of other acid gases. Dissolved oxygen<br />

can cause severe corrosion at extremely low<br />

concentration and evaporate large amounts of metal.<br />

Pitting is usually shallow, flat-bottomed, and broad-based<br />

with the tendency of one pit to combine with another. Pit<br />

shape characteristics may include sharp edges and steep<br />

sides if accompanied by CO 2 or broad, smooth craters<br />

with beveled edges if accompanied by H 2 S. Corrosion<br />

rates increase with increased concentrations of dissolved<br />

oxygen.<br />

Figure 24<br />

Figures 24 and 25 are examples of oxygen<br />

enhanced corrosion. The coupling sample on the left is<br />

an example of the effects of oxygen enhanced CO 2<br />

corrosion (left), H 2 S corrosion (middle), and chloride<br />

corrosion (right) while the rod samples in Figure 25 show<br />

the effects of oxygen enhanced CO 2 corrosion near the<br />

upset (left) and CO 2 corrosion on the rod body (right).<br />

Figure 25<br />

Scale Corrosion<br />

Scales such as barium sulfate, calcium<br />

carbonate, calcium sulfate, iron carbonate, iron oxide<br />

(rust), iron sulfide, and strontium sulfate should be prevented from forming on sucker rods. Although scale on a<br />

sucker rod slows down the corrosion penetration rate, it also reduces the effectiveness of chemical inhibitors.<br />

Severe localized corrosion in the form of pitting results any time the scale is cracked by a bending movement or<br />

removed by abrasion.<br />

12


Stray Current Corrosion<br />

Rarely seen in most producing wells, stray current corrosion refers to the induced, or stray, electrical<br />

currents that flow to or from the rod string. Stray current corrosion can be caused by grounding electrical<br />

equipment to the well casing or from nearby cathodic protection systems. Arcs originating from sucker rods<br />

leave a deep, irregular shaped pit with smooth sides, sharp edges and a small cone in the base of the pit. Arcs<br />

originating from the tubing leave deep pits with smooth sides and sharp edges that are random in dimension<br />

and irregular in shape. Stray current corrosion pits are usually singular and isolated in a row down one side of<br />

the sucker rod near the upsets.<br />

Manufacturing Defects<br />

<strong>Failure</strong>s due to manufacturing defects are rare<br />

and seldom occur. Manufacturing defects are easily<br />

recognized and it is important that you understand what<br />

these defects look like if you are to file accurate claims for<br />

warranty reimbursement. No manufacturer is excluded<br />

from the possibility of defects in material or workmanship<br />

and the following failure examples include defects from all<br />

manufactures.<br />

Figure 26<br />

Figure 26 is an example of mill defects. Mill<br />

defects occur along one side of the rod body and these<br />

discontinuities normally have longitudinally tapered, sharp<br />

“V” shaped bottoms with indications of the longitudinal seam in the base. The example on the far left is an<br />

example of a sliver. The rod body third from the left is also an example of a sliver. When fishing the rod failure,<br />

the sliver folded against the fracture surface. The rod body second from right is an example of a scab. A sliver<br />

is a small loose or torn segment and a scab is a large loose or torn segment of material longitudinally rolled into<br />

the surface of the bar. One end of the sliver or scab is normally metallurgically bonded into the rod body while<br />

the remaining end is rolled into the surface and physically attached. Fatigue failures, which result from slivers or<br />

scabs, will have a piece of loose material protruding over the fatigue portion of the fracture surface. The rod<br />

body second from the left is an example of rolled-in-scale. Rolled-in-scale is a surface discontinuity caused<br />

when scale (metal-oxide), formed during a prior heat, has not been removed prior to bar rolling. The rod body<br />

sample on the far right is an example of a rolling lap. Rolling laps are longitudinal surface discontinuities that<br />

have the appearance of a seam from rolling, with sharp corners folded over and rolled into the bar surface<br />

without metallurgical bonding.<br />

Figure 27 is an example of forging defects. The<br />

fracture begins internally below a forging crack in the<br />

upset area and is brittle or granular in appearance. A<br />

crack initiation site may or may not be visible and a<br />

fatigue portion may not be present on the fatigue fracture<br />

surface. The examples on the left and in the middle occur<br />

as a result of low forging temperatures. The example on<br />

the left is a failure from cold-shut and the example in the<br />

middle is a failure from a forging crack. The fracture on<br />

the right is a failure caused by a subsurface longitudinal<br />

seam located near the end of the raw bar. During the<br />

forging process the orientation of this discontinuity was<br />

changed transversely.<br />

Figure 27<br />

13


Figure 29 is an example of processing<br />

defects. The lower example is a casehardened<br />

sucker rod and the upper example is a coupling that<br />

has been processed through a grinding operation to<br />

reduce the diameter. In both examples, a difference<br />

in the material hardness has resulted in preferential<br />

corrosion attack.<br />

Figure 30 is an example of a mill defect and<br />

a machining defect. The lower example is a failure<br />

due to a large, internal, nonmetallic inclusion in the<br />

pin. The fracture began internally and is brittle or<br />

granular in appearance. A crack initiation site may or<br />

may not be visible and a fatigue portion may not be<br />

present on the fatigue fracture surface. The upper<br />

example is from rolling the pin threads twice. Rolling<br />

twice has flattened the pin thread crest and will not<br />

be capable of achieving the correct friction load<br />

required for makeup.<br />

Figure 29<br />

Figure 30<br />

Your initial investment in sucker rods is<br />

substantial. Moreover, the cost related to replacing<br />

damaged sucker rods generally out weights the<br />

original cost of the new rod string. Protecting your<br />

investment and getting the maximum service life out<br />

of your rod string just makes good sense. It is<br />

important to diagnose rod failures accurately and to implement corrective action measures to prevent future<br />

failure occurrences. This photo essay is intended for use as a reference guide in sucker rod failure analysis. It<br />

explains how rod failures occur and expounds methods for identifying the characteristics of the several failure<br />

mechanisms. Where sucker rod failures are concerned, there are no absolutes and no two fractures look<br />

exactly alike in appearance. But, by recognizing the visual clues and identifying characteristics of the different<br />

failure mechanisms, corrective action measures can be taken to prevent sucker rod failures, thus allowing the<br />

operator to produce marginally profitable wells more cost effectively.<br />

14


For further information please contact:<br />

<strong>Alberta</strong> <strong>Oil</strong> <strong>Tool</strong><br />

9530 – 60 th Avenue<br />

Edmonton <strong>Alberta</strong> T6E 0C1<br />

Phone: (780) 434-8566<br />

Fax: (780) 436-4329<br />

www.albertaoiltool.com<br />

Issue Date: 10/30/01

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!